Download How do glycolytic enzymes favour cancer cell proliferation by

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

RNA-Seq wikipedia , lookup

Polyclonal B cell response wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Gene regulatory network wikipedia , lookup

Signal transduction wikipedia , lookup

Metabolism wikipedia , lookup

Biochemical cascade wikipedia , lookup

Vectors in gene therapy wikipedia , lookup

Paracrine signalling wikipedia , lookup

Secreted frizzled-related protein 1 wikipedia , lookup

Glycolysis wikipedia , lookup

Transcript
Oncogene (2015) 34, 3751–3759
© 2015 Macmillan Publishers Limited All rights reserved 0950-9232/15
www.nature.com/onc
REVIEW
How do glycolytic enzymes favour cancer cell proliferation
by nonmetabolic functions?
H Lincet1,2,3 and P Icard1,4
Cancer cells enhance their glycolysis, producing lactate, even in the presence of oxygen. Glycolysis is a series of ten metabolic
reactions catalysed by enzymes whose expression is most often increased in tumour cells. HKII and phosphoglucose isomerase (PGI)
have mainly an antiapoptotic effect; PGI and glyceraldehyde-3-phosphate dehydrogenase activate survival pathways (Akt and so
on); phosphofructokinase 1 and triose phosphate isomerase participate in cell cycle activation; aldolase promotes epithelial
mesenchymal transition; PKM2 enhances various nuclear effects such as transcription, stabilisation and so on. This review outlines
the multiple non-glycolytic roles of glycolytic enzymes, which are essential for promoting cancer cells' survival, proliferation,
chemoresistance and dissemination.
Oncogene (2015) 34, 3751–3759; doi:10.1038/onc.2014.320; published online 29 September 2014
INTRODUCTION
In normal tissue, the vast majority of nonproliferating differentiated cells use oxidative phosphorylation (OXPHOS) for ATP
production. These cells metabolise glucose to pyruvate through
glycolysis, then oxidise this pyruvate through the tricarboxylic acid
cycle, generating ATP through ATP synthase, the rate of the
production being coupled with proton transport and on oxygen
respiration.1 In contrast, rapidly proliferating tumour cells consume glucose at a higher rate compared to normal cells and part
of their glucose carbon is converted into lactate, even in oxygenrich conditions; this is referred to as the ‘Warburg effect’ or
‘aerobic glycolysis’. For ATP production, cancer cells may enhance
the β-oxidation of lipids, the oxidative use of glutamine and/or
lactate.2–4 Furthermore, amino-acid uptake (alanine, aspartate,
methionine and so on) is essential to replenish the tricarboxylic
acid.5 The Warburg effect seems to be a consequence to
uncoupling between glycolysis and tricarboxylic acid-OXPHOS
due to pyruvate dehydrogenase inactivation and lactate dehydrogenase activation, in relation to HIF-1 and c-Myc activation.6,7
Since 2011, it is widely acknowledged that metabolic reprogramming is a hallmark of cancer,8 providing tumour cells with all the
metabolites (such as serine), derived from glucose and glutamine
metabolism,9,10 they need for growth and proliferation, such as
nucleotides, macromolecules, lipids and also NAD+, NADPH,H+
cofactors. Currently, a large body of evidence supports the idea
that activated oncogenes, inactivated tumour suppressors and
transcriptional factors are linked directly or indirectly to the
cellular metabolic reprogramming, establishing a relationship
between genetic alterations and glucose metabolic phenotype.5,6,11–13 The glycolytic pathway is a series of ten metabolic
reactions catalysed by multiple enzymes or enzyme complexes
whose expression is most often increased in tumour cells. The aim
of this review is to emphasise their non-glycolytic functions such
as their nuclear role (DNA repair, transcription and so on) and their
1
implications in many other functions, such as apoptosis,
detoxification, cell cycle control, signalling pathways and so on.
Hexokinases
In the cytosol, glucose (or fructose) is phosphorylated by
hexokinases (HK) (glucose kinase or fructose kinase) to glucose6-phosphate (G6P). HK catalye the first irreversible reaction of
glycolysis. Among the four mammalian HK isoenzymes (HKI to
HKIV), HKII is highly expressed in many cancers.14,15 This
predominant isoform has no regulatory site,16 a characteristic
favouring resistance to the Pasteur effect. HKII involved in the
diversion of glucose towards glycolysis or pentose phosphate
pathway (PPP) that sustain anabolic pathways such as glycerol and
serine or nucleotide biosynthesis, respectively.17 Moreover, HKII is
bound to outer mitochondrial membrane porins (voltage-gated
anion channels) to facilitate access for newly synthesised ATP for
phosphorylation glucose.18 This binding of HKII to the mitochondria decreases the negative feedback of G6P on HK, increasing
glucose metabolism via glycolysis in tumours.19
The interaction of HKII with the voltage-gated anion channels
depends on glucose uptake. Glucose deprivation prevents this
interaction in the outer mitochondrial membrane, which modifies
mitochondrial membrane potential leading to apoptosis through
the release of apoptogenic molecules.20,21 These mechanisms
could be implied in BH3-only molecules, in particular Bad and
Bid,22–24 which may link glycolysis and apoptosis in response to
glucose uptake.25
Numerous studies indicate that HKII overexpression has a
central role in the development of cancers and is associated with
oncogenic transformation and poor prognosis.26–29 This upregulation has been demonstrated to be induced by HIF-1α in certain
tumours23 and is also believed to be due to the c-Myc
oncogene.30,31 Furthermore, mRNA and protein expression levels
of HKII are regulated by microRNAs (miRs). The overexpression of
Locally Aggressive Cancer Biology and Therapy Unit (BioTICLA), Caen, France; 2Normandie University, Caen, France; 3François-Baclesse Centre for Cancer, Caen, France and
Ecole Polytechnique, Laboratoire d’Informatique, Palaiseau, France. Correspondence: Dr H Lincet, Locally Aggressive Cancer Biology and Therapy Unit (BioTICLA), EA 4656 Centre
François-Baclesse, Av Général Harris, Caen F-14032, France.
E-mail: [email protected]
Received 27 June 2014; revised 23 August 2014; accepted 23 August 2014; published online 29 September 2014
4
Glycolytic enzymes in cancer cell proliferation
H Lincet and P Icard
3752
Pyr
F6P
G6P
PGI
Apaf-1gene
PGI/AMF
Caspase-9 gene
Akt and Erk 1/2 pathways
NF-κB
ZNF217
EMT
DHAP
F1,6BP ALDO A
G6P
TPI
Pyr
GA3P
Kinases (Akt, Erk2)
EMT
E-Cadherin, β-catenin
Transcription of genes involved in S phase
Fibronectin,Vimentin
Stabilization
Protection DNA damage
Figure 1. Schematic representation of non-glycolytic functions of PGI and aldolase A (ALDO A). (a) PGI acts as cytokine outside the cell with
properties of the autocrine motility factor (AMF) referred to as PGI/AMF. It indirectly regulates expression of Apaf-1 and caspase-9 genes,
blocking the formation of apoptosome. PGI/AMF could induce cell proliferation and reduce apoptosis through PI3K/Akt and Erk1/2 pathways.
Expression of PGI/AMF is implied in the epithelial-to-mesenchymal transition (EMT) through the upregulation of NF-κB factor and ZNF217
factor. (b) ALDO A may be phosphorylated by Akt or Erk 2 kinases then may translocate in the nucleus. In this compartment, ALDO A is
involved in the regulation of transcription of genes implied in cell cycle progression, but also in the stabilisation of transcripts by linking with
AT-rich DNA sequences, and in the protection of DNA in response to DNA damage. In cytosol, ALDO A is implied in the EMT through
modifications in the expression of markers of this process.
miR-143, tumour-suppressive miRNA or anti-oncomiR, inhibits
both mRNA and protein expression of HKII, and represses
metabolism in cancer cells.32,33 In prostatic and breast tumours,
an inverse correlation has been found between miR-143 expression and HKII level, such as a decrease in miR-143 expression
linked to an increase in HKII expression.32 This miR-143 has been
found to be under indirect control of miR-155,33 which is an
oncomiR-repressing tumour suppressor gene.34
Phosphohexose isomerase (phosphoglucose isomerase, PGI)
PGI (also referred to as phosphohexose isomerase or G6P
isomerase) is a second glycolytic enzyme that catalyses the
isomerisation of G6P into fructose-6-phosphate. PGI has other
biological roles when secreted as an extracellular cytokine with
properties that include autocrine motility factor (AMF), also
referred to as PGI/AMF.35,36 Overexpressed secretion of PGI/AMF
has been observed in the serum and urine of patients with
gastrointestinal, kidney and breast cancers.36,37 PGI/AMF induces
the transformation and survival of NIH-3T3 fibroblasts,38 promotes
cell motility39 and is associated with cancer progression and
metastasis,40,41 notably through ERK activation that produces IL-8
in response to PGI/AMF.42 Moreover, PGI/AMF inhibits expression
of Apaf-1 and caspase-9 genes, and indirectly regulates the
formation of the apoptosome that induces tumour resistance to
apoptosis. The relationship between PGI/AMF and inhibition of
Apaf-1 and caspase-9 expressions remains to be studied43
(Figure 1a). PGI/AMF could also reduce apoptosis through PI3K/
Akt signalling pathway activation. This action seems a consequence of an interaction between PGI/AMF and HER2, inducing
Oncogene (2015) 3751 – 3759
HER2 phosphorylation. This pathway activates phosphoinositide3-kinase and mitogen-activated protein kinase signalling leading
Akt activation.38,44 This activation is generally associated with
increased tumour progression, tumour cell invasiveness and antiapoptosis.45 Downregulation of PGI/AMF decreases the expression
of Akt and Erk1/246 and induces epithelial–mesenchymal transition (EMT), hence reducing malignancy.47 Thus, PGI/AMF is
implied in the regulation of EMT, a phenomenon that is associated
with acquisition of invasive phenotype by cancer cells.48 PGI/AMF
upregulates the NF-κB transcription factor and ZNF217 factor,
which contribute towards the induction of EMT, and downregulates the mi-R-200 family, which is involved in the inhibition
of this process49,50 (Figure 1a).
Phosphofructokinase 1 (PFK1)
PFK1 converts fructose-6-phosphate into fructose-1,6-bisphosphate. This is the second irreversible reaction of the glycolytic
pathway and a major glycolysis checkpoint.11 PFK1 is stimulated
by ADP/AMP, whereas citrate, long-chain fatty acids, lactate and
ATP act as strong inhibitors, providing negative feedback for
glycolytic rate,51,52 (for a review, refer to Ros and Schulze53).
Moreover, in response to hypoxia, the O-linked β-N-acetylglucosamine (O-GlcNAc) post-translational modification is dynamically
induced at Ser529 of PFK1, hence inhibiting its activity.54 This
regulation reorients a great share of glucose flux through the PPP,
and results in an increase in NADPH, which is an essential cofactor
for lipogenesis and also contributes towards GSH, reducing power
regeneration and conferring cancer cells with a growth advantage.
© 2015 Macmillan Publishers Limited
Glycolytic enzymes in cancer cell proliferation
H Lincet and P Icard
The most potent allosteric activator of PFK1 is fructose-2,6bisphosphate (F2,6BP), which induces PFK1 activity even in the
presence of ATP, suggesting that it could have a role in PFK1
upregulated activity in many tumours.55 The synthesis and
degradation of F2,6BP depend upon 6-phosphofructo-2-kinase
(PFK2)/F2,6BPase (PFK2), which includes both kinase and phosphatase activities. There are four mammalian PFKFB isoenzymes
(PFKFB 1 to 4).56 The PFKFB3 isoenzyme has the highest kinase/
phosphatase activity ratio (740/1) and thus leads to elevated
F2,6BP levels, which in turn sustain high glycolytic rates. In
contrast, it has been shown that F2,6BPase overexpression induces
a decrease in the rate of F2,6BP,thereby blocking glycolysis. F2,6BP
expression is high in a number of aggressive cancers (e.g. breast,
colon and ovarian)57 and is induced by hypoxia in cultured human
colon adenocarcinoma cells.58 PFKFB3 overexpressed activity
induces the expression of several cell cycle regulators such as
phosphatase, Cdc 25C, and cyclin D3 and represses the cell cycle
inhibitor p27, all of which contribute towards an increase in CDK1
activity.59,60
Aldolase (ALDO) and Triose phosphate isomerase (TPI)
Fructose-bisphosphate ALDO catalyzes the reversible reaction of
fructose-1,6-bisphosphate to glyceraldehyde-3-phosphate and
dihydroxyacetone phosphate. In human tissue, three ALDO
isozymes are expressed in a tissue-dependent manner: ALDO A
(mainly in muscles), ALDO B (mainly in liver) and ALDO C
(expressed in neuronal tissue).61 ALDO A is the most commonly
expressed in tumour tissue62 and is overexpressed in various
cancers such as squamous cell lung cancer63,64 and hepatocellular
carcinoma.65
In addition to its glycolytic function, ALDO A is involved in
several functions, which are distinct from its glycolytic role, such as
signal transduction,66 vesicle trafficking,67 and cell motility.68
Several studies have demonstrated that ALDO A interacts with the
F-actin, which is involved in cell division during cytokinesis. Thus,
ALDO A knockdown reduces cell proliferation but is apparently
not effective on glycolytic flux, and increases multinucleation.69
Recently, ALDO A overexpression has been shown to promote
EMT and cell migration by decreasing E-cadherin and β-catenin,
and by concomitantly increasing fibronectin and vimentin, all
processes that favour tumorigenicity63 (Figure 1b). ALDO A has
also been found in the nuclear localisation of many types of
tumour,70,71 whereas decreased proliferation has been shown to
induce a reduction in ALDO in the nucleus.69 The mechanisms
involved in the subcellular localisation of ALDO A are not yet
clearly known and it might act as a cofactor favouring
proliferation. Its transport towards the nucleus may be regulated
by phosphorylation as described in vitro by PKCμ and ERK2.72 The
inhibition of kinases, such as Akt and PKC that are frequently
overexpressed in carcinogenesis73 have been shown to alter ALDO
A nuclear localisation.74 In the nuclear compartment, ALDO A
associates with nucleic acids, particularly with AT-rich DNA
sequences.71,75 This association may have a role in the stabilisation
of transcripts76 and could also be involved in agents such as
ultraviolet radiation77 (Figure 1b). ALDO A seems to be related
with the activation of transcription of certain genes involved in S
phase and may also be correlated with proliferative cell activity
(for example, DNA replication)69,74,75 (Figure 1b).
TPI is an homodimeric enzyme that converts dihydroxyacetone
phosphate into glyceraldehyde-3-phosphate, and only the latter
can complete glycolysis and thus prevent the accumulation of
dihydroxyacetone phosphate.78 TPI has been found in increased
levels in lung cancers,79 squamous cell lung carcinomas80 and
urinary cancers.81 It has also been shown to be upregulated in
chemoresistant ovarian carcinoma compared to sensitive parental
cell lines.82 In fine, activation of this enzyme is responsible for ATP
production by glycolysis, whereas its inhibition or slow-down
© 2015 Macmillan Publishers Limited
orients dihydroxyacetone phosphate towards the PPP,83 which is a
key source of reduced NADPH, a cofactor for anabolic pathways
(for example, fatty acid biosynthesis) and in maintaining the redox
balance.84 Moreover, PPP produces ribose and is involved in the
transcription of gene expression during stress conditions.85 Thus,
TPI regulates the distribution of metabolites between glycolysis or
PPP, by coordinating NADPH,H+ and ATP production with redox
metabolism depending on reactive oxygen species production.
Gruning et al. demonstrated that an accumulation of phosphoenolpyruvate inhibits TPI activity,83,86 which in turn promotes the
PPP pathway, protecting cells against oxidants and, consequently,
preventing reactive oxygen species accumulation. In contrast, the
loss of TPI catalytic activity is a consequence of phosphorylation of
this glycolytic enzyme by the cyclin A/Cdk2 complex, resulting in
disrupted energy production in etoposide-treated HeLa cells,
rendering these cells sensitive to apoptosis.87 Interestingly,
upregulation of TPI expression could partially reverse the
multidrug-resistant phenotype of SGC7901/VCR, suggesting that
TPI may be an anti-drug-resistant agent in gastric cancer and a
target candidate for developing novel therapeutics for better
treatment of gastric cancer.88
Glyceraldehyde-3-phosphate dehydrogenase (GAPDH)
GAPDH is one of most the important enzymes involved in cell
energy metabolism as it produces glycerate-1,3-biphosphate and
NADH,H+, an essential cofactor required for ATP production by
OXPHOS, and/or to sustain lactate dehydrogenase-5 activity. It is
noteworthy that lactate dehydrogenase-5 activity, which is
stimulated by HIF13 and is crucial to regenerate NAD+, required
for GAPDH activity. The GAPDH gene is classically used as a
housekeeping gene, but it is known to be overexpressed in many
tumours compared to in normal tissue, and correlated with poor
prognosis or tumour aggressiveness (for example breast, colorectal, glioma, lung, melanoma ovarian, pancreatic, prostatic and
renal cancer) (for a review, refer to Guo et al.89), and increased
drug resistance.90
Various non-glycolytic functions of GAPDH have been reported in
cancer, such as involvement in cell death,91–94 DNA repair,95,96
immunity97 (for a review, refer to Tristan et al.98) and cancer cell
senescence.99 These functions are dependent on the localisation of
the enzyme (mainly located in cytoplasm, but also in nucleus,
mitochondria and vesicular fractions).98,100,101 Exposure to
various stresses (for example, DNA damage, S-nitrosylation),92
(for a review, refer to Huang et al.102) modification of conformation
(homo-oligomerisation, tetrameric structure) and post-translational
modifications of GAPDH such as acetylation103–106 and
phosphorylation107 influence the localisation and function of this
enzyme. It has been reported that GAPDH overexpression is
associated with cell proliferation via its effect on cyclin B-cdk1
activity.108 Moreover, in the case of abundant glucose, GAPDH links
its substrate, glyceraldehyde-3-phosphate, liberating the small
GTPase Rheb that is recognised as an important positive regulator
of mTORC1,109 a crucial pathway regulating many cellular
responses to growth factors73 (Figure 2).
GAPDH seems to have controversial action on apoptosis.
Certain studies have reported a proapoptotic function92,105 when
GAPDH is located in the mitochondria and associated with the
voltage-dependent anion channel 1 to induce apoptosis.94
Contrariwise, others have described a protective role and
its implication in tumour progression110 (for a review, refer to
Colell et al.91).
Phosphoglycerate mutase (PGAM)
PGAM catalyses the interconversion of 3-phosphoglycerate (3-PG)
and 2-phosphoglycerate (2-PG), corresponding to the eighth
reaction of glycolysis. PGAM activity is commonly upregulated in
many human cancers80,111,112 and its inhibition is lethal to cancer
Oncogene (2015) 3751 – 3759
3753
Glycolytic enzymes in cancer cell proliferation
H Lincet and P Icard
3754
Pyr
1,3-BPG
GA3P
G6P
GAPDH
Inactive
mTORC1
Rheb
Rheb
Active
mTORC1
Autophagy
(Mitophagy)
TFEB
SREBP
Fatty acid
Synthesis
HIF-1α
STAT3
S6K1
PPAR
c-myc
PDK1
Cholesterol
synthesis
Mcl-1
(HK, GAPDH, LDH)
Warburg effect
Bad
HnRNPs
PPP
PGK1 PKM2
Survival
Alternatif splicing
PDH
Glycolytic enzymes
Warburg effect
Glycolysis
« Energy»
MEK5/ERK5
pathway
Proliferation
Angiogenesis
Bad
Figure 2. Multifaceted functions of GAPDH. GAPDH is a crucial factor linking metabolism and tumour growth. When glycolysis is stimulated,
GAPDH releases, in cytosol, the small GTPase Rheb that is a positive regulator of mTORC1, a crucial signalling pathway for stimulating cell
growth. The Rheb-mTORC1 complex regulates several phenotypic characteristics of cancer cells supporting proliferation and survival:
metabolic reprogramming (Warburg effect) furnishing energy to anabolic biosynthesis sustaining division (PPP, fatty and cholesterol
synthesis); apoptotic protection and various mechanisms such as autophagy, mitophagy and angiogenesis, which sustain cell metabolism.
Nucleotides
biosynthesis
P
PGAM1
6-P-gluconate
Active
SIRT1
AcAcAc
P
GPDH
PGAM1 ACTIVE
2-PG
3-PG
G6P
PGDH
Serine
Pyr
Nucleotides
biosynthesis
Protein
biosynthesis
Cancer cell proliferation
Tumor growth
Figure 3. (a) PGAM regulation and allosteric action of PGAM1. PGAM1 regulates glycolytic flux and serine synthesis. Acetylation of PGAM1
stimulates its glycolytic activity and enhances glycolysis. In return, the 3-PG level is decreased, abrogating its inhibition to glucose-6phosphate dehydrogenase, while 2-PG level is increased. This molecule favours serine synthesis by activating phosphoglycerate
dehydrogenase. Of note, SIRT1 level leads to deacetylation of PGAM1 and decreases its enzymatic activity. (b) Control of PGAM1 activation.
Tyrosine 26 phosphorylation of PGAM1 (due to growth factors such as fibroblast growth factor) contributes towards favouring 2,3bisphosphoglycerate (2,3-PG) binding and stabilisation on PGAM1. In this setting, PGAM1 is more active and promotes cancer cell
proliferation and tumour growth.
cells in culture.113 PGAM plays a crucial role in coordinating
glycolysis and biosynthesis of pathways to promote cancer cell
proliferation, particularly in hypoxia. Indeed, under hypoxic
conditions, the expression of the PGAM gene is induced, resulting
in increased protein expression and concomitant elevation of
Oncogene (2015) 3751 – 3759
PGAM enzymatic activity.114,115 This induction might contribute
towards the regulation of glycolytic flux and cell adaptation
hypoxia.116 PGAM is a critical regulatory step in glycolysis as PGAM
inhibition by epoxide inhibitors is lethal to cancer cells.117
Overexpressed PGAM can maintain its substrate (3-PG) at low
© 2015 Macmillan Publishers Limited
Glycolytic enzymes in cancer cell proliferation
H Lincet and P Icard
3755
Growth factors (EGF, IL3, …)
P
P
P
S37
P
Erk
PKM2
inactive
C-myc
P
-
P
HK, GAPDH, LDH
HnRNPs
F1,6BP,
Serine
Alternative splicing
PKM2
Metabolic genes
Alanine, …
ATP,
lipids
+
Ac
P300
Lys433
PKM2
active
Ac
HIF-1
P
MEK5, HIF1α, etc.
Ac
STAT3
JMJD5
P
JMJD5/PKM2
Figure 4. Schematic representation of non-glycolytic functions of PKM2. The regulation between tetrameric and dimeric PKM2 form is an
oscillating phenomenon subject to allosteric regulation: glycolytic intermediate F1,6BP and the biosynthetic by-product serine induce the
active form of PKM2 whereas high concentration of ATP and/or of downstream biosynthetic products (amino acids and lipids) lead the
inactive dimeric form of PKM2. This form can be translocated from cytoplasm to nucleus in response to various processes such as
phosphorylation, acetylation on lysine 433 and interaction with a dioxygenase/demethylase, Jumonji C domain-containing dioxygenase
(JMJD5). In the nucleus, JMJD5/PKM2 acts as a direct transcriptional coactivator of the metabolic genes under control of HIF1α. Nuclear PKM2
dimeric also induces c-Myc expression, leading to the upregulation of glycolytic genes and of HnRNPs, which favours the alternative splicing
of the PK isoform M2. Acetylation of PKM2 on lysine 433 leads to STAT3 phosphorylation, which enhances the transcription of genes such as
mek5 and hif1-α genes, which are essential for tumour cell proliferation, migration and adhesion.
levels, abrogating its inhibition of the G6P dehydrogenase enzyme
involved in PPP. Furthermore, the increase in 2-PG activates the
phosphoglycerate dehydrogenase enzyme, which diverts the 3-PG
to serine biosynthesis, thus contributing towards the needs of
rapidly growing tumours9,118 (Figure 3a).
Acetylation and phosphorylation are post-translational modifications in cells and the majority of enzymes participating in
glycolysis are acetylated on lysine residues, including PGAM.119
Thus, PGAM1 regulates anabolic biosynthesis by controlling
intracellular levels of substrate, the 3-PG, and product, the 2-PG.
Acetylated PGAM displays enhanced activity, whereas Sirt1, a
member of the NAD+-dependent protein deacetylases, leads to
PGAM deacetylation and downregulation of mutase activity.120
This deacetylation maintains a high level of intracellular 3-PG,
which in turn reduces PPP level by inhibiting G6P dehydrogenase.
Phosphorylated PGAM (Tyr26) by fibroblast growth factor receptor
1 induces release of inhibitory Glu19 by facilitating access to the
active site and by stabilising active conformation121 (Figure 3b).
This phosphorylation is crucial for the mutase reaction, transforming 3-PG to 2-PG in glycolysis118,122 and inducting the 2-PG level,
which in turn sustains phosphoglycerate dehydrogenase. This
enzyme diverts 3-PG from glycolysis to serine synthesis and
contributes towards maintaining a low level of 3-PG in cancerous
tumours. Moreover, the phosphoenolpyruvate pyruvate kinase
(PK) substrate can phosphorylate PGAM at the catalytic His11
(H11), producing pyruvate (the same applies in the absence of
PKM2 activity)123 and 2,3-BPG, which is a cofactor for phosphorylating His11122 (Figure 3b).
© 2015 Macmillan Publishers Limited
Pyruvate kinase
PK catalyses the tenth and last reaction of glycolysis, which also is
an irreversible reaction by transferring the phosphate from
phosphoenolpyruvate to ADP to produce ATP and pyruvate.124
There are two distinct PK genes: PKL/R (pyruvate kinase, liver and
red blood cells) and PKM2 (pyruvate kinase, muscle), which
express four PK isoforms: L, R, M1 and M2,125,126 (for a review, refer
to Mazurek127). PKM1 and PKM2 are encoded by alternative
splicing of the PKM gene125,128 which is notably regulated by three
heterogeneous nuclear ribonucleoproteins (hnRNPA1, A2 and I),
themselves regulated by c-Myc.129–131 Elevated PKM2 expression
is now known as a common characteristic of all cancers132 and is
closely associated with poor prognosis in certain tumours133–136
(for a review, refer to Wong et al.132).
PK is expressed in most cells and exists as inactive monomers,
less active dimers and active tetramers. The balance between
tetrameric and dimeric PKM2 is an oscillating phenomenon
subject to allosteric regulation: it is activated by the glycolytic
intermediate F1,6BP and by the biosynthetic by-product serine
and it is inactivated by the concentration of ATP and downstream
biosynthetic products (alanine, amino acids and lipids)
increases3,127,137–140 (Figure 4). Furthermore, PK is downregulated
by phosphorylation on Tyr105 by fibroblast growth factor receptor
1, Bcr-Abl and JAK2.141,142 Although dephosphorylated PKM2
tetramers orient pyruvate towards the tricarboxylic acid cycle to
produce ATP by OXPHOS, phosphorylated PKM2 dimers induce a
bottleneck that favours a high rate of biosynthesis, such as
nucleotides, phospholipids and amino acids5,143 but also indirectly
sustain the Warburg effect by regulating gene expression.144
Oncogene (2015) 3751 – 3759
Glycolytic enzymes in cancer cell proliferation
H Lincet and P Icard
3756
Besides phosphorylation, acetylation plays a role in the regulation
of PKM2, as acetylation on Lys305 decreases PKM2 activity,145,146
(for reviews, refer to Huang et al.102 and Yang and Lu147).
The relationship between dimeric or tetrameric PKM2 state
allows the proliferating cells to regulate their needs for anabolic
and catabolic pathways and is controlled by oncogenes and
tumour suppressors.129,137,148,149 MUC1 (mucin 1) is an oncoprotein that is overexpressed in many epithelial cancers150 (for a
review, refer to Kufe151). MUC1-cytoplasmic domain (MUC1-CD) is
translocated to the cell nucleus where it directly interacts with
multiple transcriptional factors and also acts as a transcriptional
coactivator by linking transcription factors, for example, STAT3,
NF-κB, β-catenin and HIF1-α (refer to the recent review by Nath
and Mukherjee152). Indeed, it is implied in the regulation of PKM2
activity by directly associating with PKM2 to enhance its activity,
whereas tyrosine phosphorylation of MUC1-CD decreases its
activity.153
PKM2 can be translocated from cytoplasm to cell nucleus under
certain conditions such as apoptotic signals,154 interleukin-3
response,155 phosphorylation by ERK1,156 acetylation on Lys
433146 and interaction with a dioxygenase/demethylase, Jumonji
C domain-containing dioxygenase (JMJD5)157 (for a review, refer
to Yang and Lu147) (Figure 4). In this compartment, PKM2 acts as
(1) a transcriptional coactivator, such as JMJD5/PKM2, which
modulates the HIF1α-mediated transcriptional reprogramming of
metabolic genes;157 nuclear PKM2 interacts with Oct-4 to enhance
Oct-4-mediated transcription potential158 and also with phosphorylated β-catenin to induce c-Myc expression leading to the
upregulation of glycolytic genes156,159 that promote tumorigenesis; (2) a kinase protein activates STAT3 via phosphorylation on
Tyr305 with phosphoenolpyruvate as a phosphate donor144 and
enhances the transcription of genes that are essential for tumour
cell proliferation, such as mek5 and hif1-α genes160 (for a review,
refer to Demaria and Poli161) but also promotes cell migration and
adhesion in a STAT3-dependent manner (Figure 4). These effects
are due to repression of E-cadherin by induction of Snail-2
expression,162,163 upregulation of matrix MMP-2 and MMP-9
(metalloproteinases-2 and -9)162 and could be responsible for
metastatic progression (for review on MMP-2 and -9, refer to
Bauvois164). Finally, PKM2 phosphorylates histone H3 on Thr11, a
process that is required for histone H3 acetylation on Lys9; they
also remove histone deacetylase 3 (HDAC3) from CCDN1 and MYC
promoter.165 All these effect favour cell proliferation, invasion,
metastasis and epigenetic regulation of gene expression.
CONCLUSIONS
Dysfunctions in oncogenes and or tumour suppressor genes cause
modifications in various intracellular signalling pathways, which
reprogramme tumour cells metabolism to allow enhanced survival
and growth. Enhanced aerobic glycolysis has a key role in this
metabolic reprogramming. The complex process of cancer growth
needs a flexible cell metabolism, in which the nonenzymatic
functions of most enzymes of glycolysis may offer an important
contribution. Although each enzyme may have multifaceted roles,
roughly HKII and PGI have mainly an antiapoptotic effect, PGI and
GAPDH activate survival pathways (Akt and so on), PFK1 and TPI
participate in cell cycle activation, ALDO promotes EMT and PKM2
enhances various nuclear effects such as transcription, stabilisation and so on. These multifaceted roles of enzymes participate
towards linking the plastic metabolism of cancer cells with various
processes sustaining cancer cell growth and implied in tumour
relapse, while potentially contributing towards the link between
obesity with shortened life span or cancer. Further studies should
consider these functions linking glycolysis and tumour growth
which very likely will allow to develop new anticancer strategies.
Oncogene (2015) 3751 – 3759
CONFLICT OF INTEREST
The authors declare no conflict of interest.
REFERENCES
1 Hinkle PC. P/O ratios of mitochondrial oxidative phosphorylation. Biochim
Biophys Acta 2005; 1706: 1–11.
2 Deberardinis RJ, Cheng T. Q's next: the diverse functions of glutamine in
metabolism, cell biology and cancer. Oncogene 2010; 29: 313–324.
3 Porporato PE, Dhup S, Dadhich RK, Copetti T, Sonveaux P. Anticancer targets in
the glycolytic metabolism of tumors: a comprehensive review. Front Pharmacol
2011; 2: 49.
4 Nieman KM, Kenny HA, Penicka CV, Ladanyi A, Buell-Gutbrod R, Zillhardt MR
et al. Adipocytes promote ovarian cancer metastasis and provide energy for
rapid tumor growth. Nat Med 2011; 17: 1498–1503.
5 Icard P, Lincet H. A global view of the biochemical pathways involved in the
regulation of the metabolism of cancer cells. Biochim Biophys Acta 2012; 1826:
423–433.
6 Cairns RA, Harris IS, Mak TW. Regulation of cancer cell metabolism. Nat Rev
Cancer 2011; 11: 85–95.
7 Semenza GL. Regulation of metabolism by hypoxia-inducible factor 1. Cold
Spring Harb Symp Quant Biol 2011; 76: 347–353.
8 Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell 2011;
144: 646–674.
9 Locasale JW, Grassian AR, Melman T, Lyssiotis CA, Mattaini KR, Bass AJ et al.
Phosphoglycerate dehydrogenase diverts glycolytic flux and contributes to
oncogenesis. Nat Genet 2011; 31: 869–874.
10 Posemato R, Marks KM, Shaul YD, Pacold ME, Kim D, Birsoy K et al. Functional
genomics reveal that the serine synthesis pathway is essential in breast cancer.
Nature 2011; 476: 346–350.
11 Chen JQ, Russo J. Dysregulation of glucose transport, glycolysis, TCA cycle and
glutaminolysis by oncogenes and tumor suppressors in cancer cells. Biochim
Biophys Acta 2012; 1826: 370–384.
12 Jang M, Kim SS, Lee J. Cancer cell metabolism: implications for therapeutic
targets. Exp Mol Med 2013; 45: e45.
13 Kroemer G, Pouyssegur J. Tumor cell metabolism: cancer's Achilles' heel. Cancer
Cell 2008; 13: 472–482.
14 Shinohara Y, Yamamoto K, Kogure K, Ichihara J, Terada H. Steady state transcript
levels of the type II hexokinase and type 1 glucose transporter in human tumor
cell lines. Cancer Lett 1994; 82: 27–32.
15 Pedersen PL, Mathupala S, Rempel A, Geschwind JF, Ko YH. Mitochondrial bound
type II hexokinase: a key player in the growth and survival of many cancers and
an ideal prospect for therapeutic intervention. Biochim Biophys Acta 2002; 1555:
14–20.
16 Wilson JE. Isozymes of mammalian hexokinase: structure, subcellular localization
and metabolic function. J Exp Biol 2003; 206: 2049–2057.
17 Patra KC, Wang Q, Bhaskar PT, Miller L, Wang Z, Wheaton W et al. Hexokinase 2 is
required for tumor initiation and maintenance and its systemic deletion is
therapeutic in mouse models of cancer. Cancer Cell 2013; 24: 213–228.
18 Mathupala SP, Ko YH, Pedersen PL. The pivotal roles of mitochondria in cancer:
Warburg and beyond and encouraging prospects for effective therapies. Biochim
Biophys Acta 2010; 1797: 1225–1230.
19 Bustamante E, Pedersen PL. High aerobic glycolysis of rat hepatoma cells in
culture: role of mitochondrial hexokinase. Proc Natl Acad Sci USA 1977; 74:
3735–3739.
20 Gottlob K, Majewski N, Kennedy S, Kandel E, Robey RB, Hay N. Inhibition of early
apoptotic events by Akt/PKB is dependent on the first committed step of
glycolysis and mitochondrial hexokinase. Genes Dev 2001; 15: 1406–1418.
21 Ravagnan L, Roumier T, Kroemer G. Mitochondria, the killer organelles and their
weapons. J Cell Physiol 2002; 192: 131–137.
22 Majewski N, Nogueira V, Bhaskar P, Coy PE, Skeen JE, Gottlob K et al. Hexokinasemitochondria interaction mediated by Akt is required to inhibit apoptosis in the
presence or absence of Bax and Bak. Mol Cell 2004; 16: 819–830.
23 Pastorino JG, Hoek JB. Hexokinase II: the integration of energy metabolism and
control of apoptosis. Curr Med Chem 2003; 10: 1535–1551.
24 Danial NN, Gramm CF, Scorrano L, Zhang CY, Krauss S, Ranger AM et al. BAD and
glucokinase reside in a mitochondrial complex that integrates glycolysis and
apoptosis. Nature 2003; 424: 952–956.
25 Yang J, Li JH, Wang J, Zhang CY. Molecular modeling of BAD complex resided in
a mitochondrion integrating glycolysis and apoptosis. J Theor Biol 2010; 266:
231–241.
26 Min JW, Kim KI, Kim HA, Kim EK, Noh WC, Jeon HB et al. INPP4B-mediated tumor
resistance is associated with modulation of glucose metabolism via hexokinase 2
© 2015 Macmillan Publishers Limited
Glycolytic enzymes in cancer cell proliferation
H Lincet and P Icard
3757
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
regulation in laryngeal cancer cells. Biochem Biophys Res Commun 2013; 440:
137–142.
Palmieri D, Fitzgerald D, Shreeve SM, Hua E, Bronder JL, Weil RJ et al. Analyses of
resected human brain metastases of breast cancer reveal the association
between up-regulation of hexokinase 2 and poor prognosis. Mol Cancer Res
2009; 7: 1438–1445.
Kwee SA, Hernandez B, Chan O, Wong L. Choline kinase alpha and hexokinase-2
protein expression in hepatocellular carcinoma: association with survival. PLoS
ONE 2012; 7: e46591.
Chen J, Zhang S, Li Y, Tang Z, Kong W. Hexokinase 2 overexpression promotes
the proliferation and survival of laryngeal squamous cell carcinoma. Tumour Biol
2014; 35: 3743–3753.
Kim JW, Gao P, Liu YC, Semenza GL, Dang CV. Hypoxia-inducible factor 1 and
dysregulated c-Myc cooperatively induce vascular endothelial growth factor
and metabolic switches hexokinase 2 and pyruvate dehydrogenase kinase 1. Mol
Cell Biol 2007; 27: 7381–7393.
Yeung SJ, Pan J, Lee MH. Roles of p53, MYC and HIF-1 in regulating glycolysis the seventh hallmark of cancer. Cell Mol Life Sci 2008; 65: 3981–3999.
Peschiaroli A, Giacobbe A, Formosa A, Markert EK, Bongiorno-Borbone L, Levine
AJ et al. miR-143 regulates hexokinase 2 expression in cancer cells. Oncogene
2013; 32: 797–802.
Jiang S, Zhang LF, Zhang HW, Hu S, Lu MH, Liang S et al. A novel miR-155/
miR-143 cascade controls glycolysis by regulating hexokinase 2 in breast
cancer cells. EMBO J 2012; 31: 1985–1998.
Jiang S, Zhang HW, Lu MH, He XH, Li Y, Gu H et al. MicroRNA-155 functions as an
OncomiR in breast cancer by targeting the suppressor of cytokine signaling 1
gene. Cancer Res 2010; 70: 3119–3127.
Watanabe H, Takehana K, Date M, Shinozaki T, Raz A. Tumor cell autocrine
motility factor is the neuroleukin/phosphohexose isomerase polypeptide. Cancer
Res 1996; 56: 2960–2963.
Liotta LA, Mandler R, Murano G, Katz DA, Gordon RK, Chiang PK et al. Tumor cell
autocrine motility factor. Proc Natl Acad Sci USA 1986; 83: 3302–3306.
Baumann M, Kappl A, Lang T, Brand K, Siegfried W, Paterok E. The diagnostic
validity of the serum tumor marker phosphohexose isomerase (PHI) in
patients with gastrointestinal, kidney, and breast cancer. Cancer Invest 1990; 8:
351–356.
Tsutsumi S, Hogan V, Nabi IR, Raz A. Overexpression of the autocrine motility
factor/phosphoglucose isomerase induces transformation and survival of
NIH-3T3 fibroblasts. Cancer Res 2003; 63: 242–249.
Niinaka Y, Paku S, Haga A, Watanabe H, Raz A. Expression and secretion of
neuroleukin/phosphohexose isomerase/maturation factor as autocrine motility
factor by tumor cells. Cancer Res 1998; 58: 2667–2674.
Nabi IR, Watanabe H, Raz A. Autocrine motility factor and its receptor: role in cell
locomotion and metastasis. Cancer Metastasis Rev 1992; 11: 5–20.
Funasaka T, Yanagawa T, Hogan V, Raz A. Regulation of phosphoglucose isomerase/autocrine motility factor expression by hypoxia. FASEB J 2005; 19:
1422–1430.
Araki K, Shimura T, Yajima T, Tsutsumi S, Suzuki H, Okada K et al. Phosphoglucose isomerase/autocrine motility factor promotes melanoma cell migration
through ERK activation dependent on autocrine production of interleukin-8.
J Biol Chem 2009; 284: 32305–32311.
Haga A, Funasaka T, Niinaka Y, Raz A, Nagase H. Autocrine motility factor
signaling induces tumor apoptotic resistance by regulations Apaf-1 and
Caspase-9 apoptosome expression. Int J Cancer 2003; 107: 707–714.
Kho DH, Nangia-Makker P, Balan V, Hogan V, Tait L, Wang Y et al. Autocrine
motility factor promotes HER2 cleavage and signaling in breast cancer cells.
Cancer Res 2013; 73: 1411–1419.
Fruman DA, Rommel C. PI3K and cancer: lessons, challenges and opportunities.
Nat Rev Drug Discov 2014; 13: 140–156.
You ZM, Zhao L, Xia J, Wei Q, Liu YM, Liu XY et al. Down-regulation of phosphoglucose isomerase/autocrine motility factor enhances gensenoside Rh2
pharmacological action on leukemia KG1alpha cells. Asian Pac J Cancer Prev
2014; 15: 1099–1104.
Niinaka Y, Harada K, Fujimuro M, Oda M, Haga A, Hosoki M et al. Silencing of
autocrine motility factor induces mesenchymal-to-epithelial transition and
suppression of osteosarcoma pulmonary metastasis. Cancer Res 2010; 70:
9483–9493.
Thiery JP, Acloque H, Huang RY, Nieto MA. Epithelial-mesenchymal transitions in
development and disease. Cell 2009; 139: 871–890.
Vendrell JA, Thollet A, Nguyen NT, Ghayad SE, Vinot S, Bieche I et al. ZNF217 is a
marker of poor prognosis in breast cancer that drives epithelial-mesenchymal
transition and invasion. Cancer Res 2012; 72: 3593–3606.
Ahmad A, Aboukameel A, Kong D, Wang Z, Sethi S, Chen W et al. Phosphoglucose
isomerase/autocrine motility factor mediates epithelial-mesenchymal transition
regulated by miR-200 in breast cancer cells. Cancer Res 2011; 71: 3400–3409.
© 2015 Macmillan Publishers Limited
51 Jenkins CM, Yang J, Sims HF, Gross RW. Reversible high affinity inhibition of
phosphofructokinase-1 by acyl-CoA: a mechanism integrating glycolytic flux
with lipid metabolism. J Biol Chem 2011; 286: 11937–11950.
52 Costa LT, Da SD, Guimaraes CR, Zancan P, Sola-Penna M. Lactate favours
the dissociation of skeletal muscle 6-phosphofructo-1-kinase tetramers
down-regulating the enzyme and muscle glycolysis. Biochem J 2007; 408:
123–130.
53 Ros S, Schulze A. Balancing glycolytic flux: the role of 6-phosphofructo-2-kinase/
fructose 2,6-bisphosphatases in cancer metabolism. Cancer Metab 2013; 1: 8.
54 Yi W, Clark PM, Mason DE, Keenan MC, Hill C, Goddard WA III et al. Phosphofructokinase 1 glycosylation regulates cell growth and metabolism. Science 2012;
337: 975–980.
55 Van SE, Hue L, Hers HG. Fructose 2,6-bisphosphate, the probably structure of
the glucose- and glucagon-sensitive stimulator of phosphofructokinase. Biochem
J 1980; 192: 897–901.
56 Rider MH, Bertrand L, Vertommen D, Michels PA, Rousseau GG, Hue L.
6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase: head-to-head with a
bifunctional enzyme that controls glycolysis. Biochem J 2004; 381: 561–579.
57 Atsumi T, Chesney J, Metz C, Leng L, Donnelly S, Makita Z et al. High expression
of inducible 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase (iPFK-2;
PFKFB3) in human cancers. Cancer Res 2002; 62: 5881–5887.
58 Chesney J. 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase and tumor
cell glycolysis. Curr Opin Clin Nutr Metab Care 2006; 9: 535–539.
59 Yalcin A, Clem BF, Simmons A, Lane A, Nelson K, Clem AL et al. Nuclear targeting
of 6-phosphofructo-2-kinase (PFKFB3) increases proliferation via cyclindependent kinases. J Biol Chem 2009; 284: 24223–24232.
60 Almeida A, Bolanos JP, Moncada S. E3 ubiquitin ligase APC/C-Cdh1 accounts for
the Warburg effect by linking glycolysis to cell proliferation. Proc Natl Acad Sci
USA 2010; 107: 738–741.
61 Mukai T, Joh K, Arai Y, Yatsuki H, Hori K. Tissue-specific expression of rat aldolase
A mRNAs. Three molecular species differing only in the 5'-terminal sequences.
J Biol Chem 1986; 261: 3347–3354.
62 Asaka M, Kimura T, Meguro T, Kato M, Kudo M, Miyazaki T et al. Alteration of
aldolase isozymes in serum and tissues of patients with cancer and other
diseases. J Clin Lab Anal 1994; 8: 144–148.
63 Du S, Guan Z, Hao L, Song Y, Wang L, Gong L et al. Fructose-bisphosphate
aldolase a is a potential metastasis-associated marker of lung squamous cell
carcinoma and promotes lung cell tumorigenesis and migration. PLoS ONE 2014;
9: e85804.
64 Poschmann G, Sitek B, Sipos B, Ulrich A, Wiese S, Stephan C et al. Identification of
proteomic differences between squamous cell carcinoma of the lung and
bronchial epithelium. Mol Cell Proteomics 2009; 8: 1105–1116.
65 Chaerkady R, Harsha HC, Nalli A, Gucek M, Vivekanandan P, Akhtar J et al.
A quantitative proteomic approach for identification of potential biomarkers
in hepatocellular carcinoma. J Proteome Res 2008; 7: 4289–4298.
66 Kim JH, Lee S, Kim JH, Lee TG, Hirata M, Suh PG et al. Phospholipase D2 directly
interacts with aldolase via Its PH domain. Biochemistry 2002; 41: 3414–3421.
67 Kao AW, Noda Y, Johnson JH, Pessin JE, Saltiel AR. Aldolase mediates the
association of F-actin with the insulin-responsive glucose transporter GLUT4.
J Biol Chem 1999; 274: 17742–17747.
68 Buscaglia CA, Penesetti D, Tao M, Nussenzweig V. Characterization of an
aldolase-binding site in the Wiskott-Aldrich syndrome protein. J Biol Chem 2006;
281: 1324–1331.
69 Ritterson LC, Tolan DR. Targeting of several glycolytic enzymes using RNA
interference reveals aldolase affects cancer cell proliferation through a nonglycolytic mechanism. J Biol Chem 2012; 287: 42554–42563.
70 Saez DE, Slebe JC. Subcellular localization of aldolase B. J Cell Biochem 2000; 78:
62–72.
71 Satou W, Tanimoto H, Ukekawa R, Fujii M, Ayusawa D. Amplification of nuclear
aldolase A in mouse cell mutants resistant to Hoechst 33342. Biochem Biophys
Res Commun 2004; 315: 845–849.
72 Gao ZH, Metherall J, Virshup DM. Identification of casein kinase I substrates by
in vitro expression cloning screening. Biochem Biophys Res Commun 2000; 268:
562–566.
73 Laplante M, Sabatini DM. Regulation of mTORC1 and its impact on gene
expression at a glance. J Cell Sci 2013; 126: 1713–1719.
74 Mamczur P, Gamian A, Kolodziej J, Dziegiel P, Rakus D. Nuclear localization of
aldolase A correlates with cell proliferation. Biochim Biophys Acta 2013; 1833:
2812–2822.
75 Ronai Z, Robinson R, Rutberg S, Lazarus P, Sardana M. Aldolase-DNA interactions
in a SEWA cell system. Biochim Biophys Acta 1992; 1130: 20–28.
76 Canete-Soler R, Reddy KS, Tolan DR, Zhai J. Aldolases a and C are ribonucleolytic
components of a neuronal complex that regulates the stability of the lightneurofilament mRNA. J Neurosci 2005; 25: 4353–4364.
Oncogene (2015) 3751 – 3759
Glycolytic enzymes in cancer cell proliferation
H Lincet and P Icard
3758
77 Lu J, Suzuki T, Satoh M, Chen S, Tomonaga T, Nomura F et al. Involvement of
aldolase A in X-ray resistance of human HeLa and UV(r)-1 cells. Biochem Biophys
Res Commun 2008; 369: 948–952.
78 Wierenga RK, Kapetaniou EG, Venkatesan R. Triosephosphate isomerase: a highly
evolved biocatalyst. Cell Mol Life Sci 2010; 67: 3961–3982.
79 Chen G, Gharib TG, Huang CC, Thomas DG, Shedden KA, Taylor JM et al. Proteomic analysis of lung adenocarcinoma: identification of a highly expressed set
of proteins in tumors. Clin Cancer Res 2002; 8: 2298–2305.
80 Li C, Xiao Z, Chen Z, Zhang X, Li J, Wu X et al. Proteome analysis of human lung
squamous carcinoma. Proteomics 2006; 6: 547–558.
81 Unwin RD, Craven RA, Harnden P, Hanrahan S, Totty N, Knowles M et al.
Proteomic changes in renal cancer and co-ordinate demonstration of both the
glycolytic and mitochondrial aspects of the Warburg effect. Proteomics 2003; 3:
1620–1632.
82 Di MM, Marcone S, Cicchillitti L, Della CA, Ferlini C, Scambia G et al.
Glycoproteomics of paclitaxel resistance in human epithelial ovarian cancer cell
lines: towards the identification of putative biomarkers. J Proteomics 2010; 73:
879–898.
83 Gruning NM, Du D, Keller MA, Luisi BF, Ralser M. Inhibition of triosephosphate
isomerase by phosphoenolpyruvate in the feedback-regulation of glycolysis.
Open Biol 2014; 4: 130232.
84 Ying W. NAD+/NADH and NADP+/NADPH in cellular functions and cell death:
regulation and biological consequences. Antioxid Redox Signal 2008; 10:
179–206.
85 Kruger A, Gruning NM, Wamelink MM, Kerick M, Kirpy A, Parkhomchuk D et al.
The pentose phosphate pathway is a metabolic redox sensor and regulates
transcription during the antioxidant response. Antioxid Redox Signal 2011; 15:
311–324.
86 Gruning NM, Rinnerthaler M, Bluemlein K, Mulleder M, Wamelink MM, Lehrach H
et al. Pyruvate kinase triggers a metabolic feedback loop that controls redox
metabolism in respiring cells. Cell Metab 2011; 14: 415–427.
87 Lee WH, Choi JS, Byun MR, Koo KT, Shin S, Lee SK et al. Functional inactivation of
triosephosphate isomerase through phosphorylation during etoposide-induced
apoptosis in HeLa cells: potential role of Cdk2. Toxicology 2010; 278: 224–228.
88 Wang X, Lu Y, Yang J, Shi Y, Lan M, Liu Z et al. Identification of triosephosphate
isomerase as an anti-drug resistance agent in human gastric cancer cells using
functional proteomic analysis. J Cancer Res Clin Oncol 2008; 134: 995–1003.
89 Guo C, Liu S, Sun MZ. Novel insight into the role of GAPDH playing in tumor. Clin
Transl Oncol 2013; 15: 167–172.
90 Demarse NA, Ponnusamy S, Spicer EK, Apohan E, Baatz JE, Ogretmen B et al.
Direct binding of glyceraldehyde 3-phosphate dehydrogenase to telomeric DNA
protects telomeres against chemotherapy-induced rapid degradation. J Mol Biol
2009; 394: 789–803.
91 Colell A, Green DR, Ricci JE. Novel roles for GAPDH in cell death and carcinogenesis. Cell Death Differ 2009; 16: 1573–1581.
92 Hara MR, Agrawal N, Kim SF, Cascio MB, Fujimuro M, Ozeki Y et al. S-nitrosylated
GAPDH initiates apoptotic cell death by nuclear translocation following Siah1
binding. Nat Cell Biol 2005; 7: 665–674.
93 Huang J, Hao L, Xiong N, Cao X, Liang Z, Sun S et al. Involvement of
glyceraldehyde-3-phosphate dehydrogenase in rotenone-induced cell apoptosis: relevance to protein misfolding and aggregation. Brain Res 2009; 1279: 1–8.
94 Tarze A, Deniaud A, Le BM, Maillier E, Molle D, Larochette N et al. GAPDH, a novel
regulator of the pro-apoptotic mitochondrial membrane permeabilization.
Oncogene 2007; 26: 2606–2620.
95 Krynetski EY, Krynetskaia NF, Bianchi ME, Evans WE. A nuclear protein complex
containing high mobility group proteins B1 and B2, heat shock cognate protein
70, ERp60, and glyceraldehyde-3-phosphate dehydrogenase is involved in the
cytotoxic response to DNA modified by incorporation of anticancer nucleoside
analogues. Cancer Res 2003; 63: 100–106.
96 Meyer-Siegler K, Rahman-Mansur N, Wurzer JC, Sirover MA. Proliferative
dependent regulation of the glyceraldehyde-3-phosphate dehydrogenase/uracil
DNA glycosylase gene in human cells. Carcinogenesis 1992; 13: 2127–2132.
97 Chang CH, Curtis JD, Maggi LB Jr, Faubert B, Villarino AV, O'Sullivan D et al.
Posttranscriptional control of T cell effector function by aerobic glycolysis. Cell
2013; 153: 1239–1251.
98 Tristan C, Shahani N, Sedlak TW, Sawa A. The diverse functions of GAPDH: views
from different subcellular compartments. Cell Signal 2011; 23: 317–323.
99 Nicholls C, Pinto AR, Li H, Li L, Wang L, Simpson R et al. Glyceraldehyde-3phosphate dehydrogenase (GAPDH) induces cancer cell senescence by interacting with telomerase RNA component. Proc Natl Acad Sci USA 2012; 109:
13308–13313.
100 Krasnov GS, Dmitriev AA, Snezhkina AV, Kudryavtseva AV. Deregulation of glycolysis in cancer: glyceraldehyde-3-phosphate dehydrogenase as a therapeutic target.
Expert Opin Ther Targets 2013; 17: 681–693.
Oncogene (2015) 3751 – 3759
101 Sirover MA. Subcellular dynamics of multifunctional protein regulation:
mechanisms of GAPDH intracellular translocation. J Cell Biochem 2012; 113:
2193–2200.
102 Huang W, Wang Z, Lei QY. Acetylation control of metabolic enzymes in cancer:
an updated version. Acta Biochim Biophys Sin (Shanghai) 2014; 46: 204–213.
103 Joo HY, Woo SR, Shen YN, Yun MY, Shin HJ, Park ER et al. SIRT1 interacts with and
protects glyceraldehyde-3-phosphate dehydrogenase (GAPDH) from nuclear
translocation: implications for cell survival after irradiation. Biochem Biophys
Res Commun 2012; 424: 681–686.
104 Li T, Liu M, Feng X, Wang Z, Das I, Xu Y et al. Glyceraldehyde-3-phosphate
dehydrogenase is activated by lysine 254 acetylation in response to
glucose signal. J Biol Chem 2014; 289: 3775–3785.
105 Sen N, Hara MR, Kornberg MD, Cascio MB, Bae BI, Shahani N et al. Nitric
oxide-induced nuclear GAPDH activates p300/CBP and mediates apoptosis.
Nat Cell Biol 2008; 10: 866–873.
106 Ventura M, Mateo F, Serratosa J, Salaet I, Carujo S, Bachs O et al. Nuclear
translocation of glyceraldehyde-3-phosphate dehydrogenase is regulated by
acetylation. Int J Biochem Cell Biol 2010; 42: 1672–1680.
107 Huang Q, Lan F, Zheng Z, Xie F, Han J, Dong L et al. Akt2 kinase suppresses
glyceraldehyde-3-phosphate dehydrogenase (GAPDH)-mediated apoptosis in
ovarian cancer cells via phosphorylating GAPDH at threonine 237 and
decreasing its nuclear translocation. J Biol Chem 2011; 286: 42211–42220.
108 Carujo S, Estanyol JM, Ejarque A, Agell N, Bachs O, Pujol MJ. Glyceraldehyde
3-phosphate dehydrogenase is a SET-binding protein and regulates cyclin
B-cdk1 activity. Oncogene 2006; 25: 4033–4042.
109 Inoki K, Li Y, Xu T, Guan KL. Rheb GTPase is a direct target of TSC2 GAP activity
and regulates mTOR signaling. Genes Dev 2003; 17: 1829–1834.
110 Colell A, Ricci JE, Tait S, Milasta S, Maurer U, Bouchier-Hayes L et al. GAPDH and
autophagy preserve survival after apoptotic cytochrome c release in the absence
of caspase activation. Cell 2007; 129: 983–997.
111 Ren F, Wu H, Lei Y, Zhang H, Liu R, Zhao Y et al. Quantitative proteomics
identification of phosphoglycerate mutase 1 as a novel therapeutic target in
hepatocellular carcinoma. Mol Cancer 2010; 9: 81–89.
112 Shen J, Wang W, Wu J, Feng B, Chen W, Wang M et al. Comparative proteomic
profiling of human bile reveals SSP411 as a novel biomarker of cholangiocarcinoma. PLoS ONE 2012; 7: e47476.
113 Evans MJ, Saghatelian A, Sorensen EJ, Cravatt BF. Target discovery in
small-molecule cell-based screens by in situ proteome reactivity profiling. Nat
Biotechnol 2005; 23: 1303–1307.
114 Hamaguchi T, Iizuka N, Tsunedomi R, Hamamoto Y, Miyamoto T, Iida M et al.
Glycolysis module activated by hypoxia-inducible factor 1alpha is related to the
aggressive phenotype of hepatocellular carcinoma. Int J Oncol 2008; 33:
725–731.
115 Takahashi Y, Takahashi S, Yoshimi T, Miura T. Hypoxia-induced expression
of phosphoglycerate mutase B in fibroblasts. Eur J Biochem 1998; 254:
497–504.
116 Sharma NK, Sethy NK, Bhargava K. Comparative proteome analysis reveals
differential regulation of glycolytic and antioxidant enzymes in cortex and
hippocampus exposed to short-term hypobaric hypoxia. J Proteomics 2013; 79:
277–298.
117 Evans MJ, Morris GM, Wu J, Olson AJ, Sorensen EJ, Cravatt BF. Mechanistic and
structural requirements for active site labeling of phosphoglycerate mutase by
spiroepoxides. Mol Biosyst 2007; 3: 495–506.
118 Hitosugi T, Zhou L, Elf S, Fan J, Kang HB, Seo JH et al. Phosphoglycerate mutase
1 coordinates glycolysis and biosynthesis to promote tumor growth. Cancer Cell
2012; 22: 585–600.
119 Zhao S, Xu W, Jiang W, Yu W, Lin Y, Zhang T et al. Regulation of cellular
metabolism by protein lysine acetylation. Science 2010; 327: 1000–1004.
120 Hallows WC, Yu W, Denu JM. Regulation of glycolytic enzyme phosphoglycerate
mutase-1 by Sirt1 protein-mediated deacetylation. J Biol Chem 2012; 287:
3850–3858.
121 Hitosugi T, Zhou L, Fan J, Elf S, Zhang L, Xie J et al. Tyr26 phosphorylation of
PGAM1 provides a metabolic advantage to tumours by stabilizing the active
conformation. Nat Commun 2013; 4: 1790.
122 Grisolia S, Cleland WW. Influence of salt, substrate, and cofactor concentrations
on the kinetic and mechanistic behavior of phosphoglycerate mutase.
Biochemistry 1968; 7: 1115–1121.
123 Vander Heiden MG, Locasale JW, Swanson KD, Sharfi H, Heffron GJ,
Amador-Noguez D et al. Evidence for an alternative glycolytic pathway in rapidly
proliferating cells. Science 2010; 329: 1492–1499.
124 Stryer L. Biochemistry. 2nd edition. W. H. Freeman and Company: New York, 1981.
125 Noguchi T, Inoue H, Tanaka T. The M1- and M2-type isozymes of rat pyruvate
kinase are produced from the same gene by alternative RNA splicing. J Biol Chem
1986; 261: 13807–13812.
© 2015 Macmillan Publishers Limited
Glycolytic enzymes in cancer cell proliferation
H Lincet and P Icard
3759
126 Noguchi T, Yamada K, Inoue H, Matsuda T, Tanaka T. The L- and R-type isozymes
of rat pyruvate kinase are produced from a single gene by use of different
promoters. J Biol Chem 1987; 262: 14366–14371.
127 Mazurek S. Pyruvate kinase type M2: a key regulator of the metabolic budget
system in tumor cells. Int J Biochem Cell Biol 2011; 43: 969–980.
128 Christofk HR, Vander Heiden MG, Harris MH, Ramanathan A, Gerszten RE, Wei R
et al. The M2 splice isoform of pyruvate kinase is important for cancer
metabolism and tumour growth. Nature 2008; 452: 230–233.
129 David CJ, Chen M, Assanah M, Canoll P, Manley JL. HnRNP proteins controlled by
c-Myc deregulate pyruvate kinase mRNA splicing in cancer. Nature 2010; 463:
364–368.
130 Chen M, Zhang J, Manley JL. Turning on a fuel switch of cancer: hnRNP proteins
regulate alternative splicing of pyruvate kinase mRNA. Cancer Res 2010; 70:
8977–8980.
131 Liu J, Wu N, Ma L, Liu M, Liu G, Zhang Y et al. Oleanolic Acid suppresses aerobic
glycolysis in cancer cells by switching pyruvate kinase type m isoforms. PLoS
ONE 2014; 9: e91606.
132 Wong N, Ojo D, Yan J, Tang D. PKM2 contributes to cancer metabolism. Cancer
Lett 2014 (in press).
133 Desai S, Ding M, Wang B, Lu Z, Zhao Q, Shaw K et al. Tissue-specific isoform
switch and DNA hypomethylation of the pyruvate kinase PKM gene in human
cancers. Oncotarget 2013; 7: 1–9.
134 Li J, Yang Z, Zou Q, Yuan Y, Li J, Liang L et al. PKM2 and ACVR 1C are prognostic
markers for poor prognosis of gallbladder cancer. Clin Transl Oncol 2014; 16:
200–207.
135 Mukherjee J, Phillips JJ, Zheng S, Wiencke J, Ronen SM, Pieper RO. Pyruvate
kinase M2 expression, but not pyruvate kinase activity, is up-regulated in a
grade-specific manner in human glioma. PLoS ONE 2013; 8: e57610.
136 Zhang X, He C, He C, Chen B, Liu Y, Kong M et al. Nuclear PKM2 expression
predicts poor prognosis in patients with esophageal squamous cell carcinoma.
Pathol Res Pract 2013; 209: 510–515.
137 Luo W, Semenza GL. Emerging roles of PKM2 in cell metabolism and cancer
progression. Trends Endocrinol Metab 2012; 23: 560–566.
138 Dhup S, Dadhich RK, Porporato PE, Sonveaux P. Multiple biological activities of
lactic Acid in cancer: influences on tumor growth, angiogenesis and metastasis.
Curr Pharm Des 2012; 18: 1319–1330.
139 Chaneton B, Hillmann P, Zheng L, Martin AC, Maddocks OD, Chokkathukalam A
et al. Serine is a natural ligand and allosteric activator of pyruvate kinase M2.
Nature 2012; 491: 458–462.
140 Israelsen WJ, Dayton TL, Davidson SM, Fiske BP, Hosios AM, Bellinger G et al.
PKM2 isoform-specific deletion reveals a differential requirement for pyruvate
kinase in tumor cells. Cell 2013; 155: 397–409.
141 Hitosugi T, Kang S, Vander Heiden MG, Chung TW, Elf S, Lythgoe K et al. Tyrosine
phosphorylation inhibits PKM2 to promote the Warburg effect and
tumor growth. Sci Signal 2009; 2: ra73.
142 Christofk HR, Vander Heiden MG, Wu N, Asara JM, Cantley LC. Pyruvate kinase
M2 is a phosphotyrosine-binding protein. Nature 2008; 452: 181–186.
143 Iqbal MA, Gupta V, Gopinath P, Mazurek S, Bamezai RN. Pyruvate kinase M2 and
cancer: an updated assessment. FEBS Lett 2014; 588: 2685–2692.
144 Gao X, Wang H, Yang JJ, Liu X, Liu ZR. Pyruvate kinase M2 regulates gene
transcription by acting as a protein kinase. Mol Cell 2012; 45: 598–609.
145 Lv L, Li D, Zhao D, Lin R, Chu Y, Zhang H et al. Acetylation targets the M2 isoform
of pyruvate kinase for degradation through chaperone-mediated autophagy
and promotes tumor growth. Mol Cell 2011; 42: 719–730.
© 2015 Macmillan Publishers Limited
146 Lv L, Xu YP, Zhao D, Li FL, Wang W, Sasaki N et al. Mitogenic and oncogenic
stimulation of K433 acetylation promotes PKM2 protein kinase activity and
nuclear localization. Mol Cell 2013; 52: 340–352.
147 Yang W, Lu Z. Regulation and function of pyruvate kinase M2 in cancer. Cancer
Lett 2013; 339: 153–158.
148 Miller DM, Thomas SD, Islam A, Muench D, Sedoris K. c-Myc and cancer
metabolism. Clin Cancer Res 2012; 18: 5546–5553.
149 Mehla K, Singh PK. MUC1: A novel metabolic master regulator. Biochim Biophys
Acta 2014; 1845: 126–135.
150 Lau SK, Weiss LM, Chu PG. Differential expression of MUC1, MUC2, and MUC5AC
in carcinomas of various sites: an immunohistochemical study. Am J Clin Pathol
2004; 122: 61–69.
151 Kufe DW. Mucins in cancer: function, prognosis and therapy. Nat Rev Cancer
2009; 9: 874–885.
152 Nath S, Mukherjee P. MUC1: a multifaceted oncoprotein with a key role in cancer
progression. Trends Mol Med 2014; 20: 332–342.
153 Kosugi M, Ahmad R, Alam M, Uchida Y, Kufe D. MUC1-C oncoprotein regulates
glycolysis and pyruvate kinase M2 activity in cancer cells. PLoS ONE 2011; 6:
e28234.
154 Stetak A, Veress R, Ovadi J, Csermely P, Keri G, Ullrich A. Nuclear translocation of
the tumor marker pyruvate kinase M2 induces programmed cell death. Cancer
Res 2007; 67: 1602–1608.
155 Hoshino A, Hirst JA, Fujii H. Regulation of cell proliferation by interleukin-3induced nuclear translocation of pyruvate kinase. J Biol Chem 2007; 282:
17706–17711.
156 Yang W, Zheng Y, Xia Y, Ji H, Chen X, Guo F et al. ERK1/2-dependent phosphorylation and nuclear translocation of PKM2 promotes the Warburg effect. Nat
Cell Biol 2012; 14: 1295–1304.
157 Wang HJ, Hsieh YJ, Cheng WC, Lin CP, Lin YS, Yang SF et al. JMJD5 regulates
PKM2 nuclear translocation and reprograms HIF-1alpha-mediated glucose
metabolism. Proc Natl Acad Sci USA 2014; 111: 279–284.
158 Lee J, Kim HK, Han YM, Kim J. Pyruvate kinase isozyme type M2 (PKM2) interacts
and cooperates with Oct-4 in regulating transcription. Int J Biochem Cell Biol
2008; 40: 1043–1054.
159 Yang W, Xia Y, Ji H, Zheng Y, Liang J, Huang W et al. Nuclear PKM2
regulates beta-catenin transactivation upon EGFR activation. Nature 2011; 480:
118–122.
160 Demaria M, Giorgi C, Lebiedzinska M, Esposito G, D'Angeli L, Bartoli A et al.
A STAT3-mediated metabolic switch is involved in tumour transformation and
STAT3 addiction. Aging (Albany NY) 2010; 2: 823–842.
161 Demaria M, Poli V. PKM2, STAT3 and HIF-1alpha: The Warburg's vicious circle.
JAKSTAT 2012; 1: 194–196.
162 Yang P, Li Z, Fu R, Wu H, Li Z. Pyruvate kinase M2 facilitates colon cancer cell
migration via the modulation of STAT3 signalling. Cell Signal 2014; 26:
1853–1862.
163 Wang LY, Liu YP, Chen LG, Chen YL, Tan L, Liu JJ et al. Pyruvate kinase M2 plays a
dual role on regulation of the EGF/EGFR signaling via E-cadherin-dependent
manner in gastric cancer cells. PLoS ONE 2013; 8: e67542.
164 Bauvois B. New facets of matrix metalloproteinases MMP-2 and MMP-9 as cell
surface transducers: outside-in signaling and relationship to tumor progression.
Biochim Biophys Acta 2012; 1825: 29–36.
165 Yang W, Xia Y, Hawke D, Li X, Liang J, Xing D et al. PKM2 phosphorylates histone
H3 and promotes gene transcription and tumorigenesis. Cell 2012; 150:
685–696.
Oncogene (2015) 3751 – 3759